You are on page 1of 11

Spectrochimica Acta Part A 84 (2011) 275–285

Contents lists available at SciVerse ScienceDirect

Spectrochimica Acta Part A: Molecular and


Biomolecular Spectroscopy
journal homepage: www.elsevier.com/locate/saa

Molecular structure and spectroscopic studies on novel complexes of


coumarin-3-carboxylic acid with Ni(II), Co(II), Zn(II) and Mn(II) ions based on
density functional theory
B.S. Creaven a,b,∗ , M. Devereux c , I. Georgieva d,∗∗ , D. Karcz a,b , M. McCann e , N. Trendafilova d ,
M. Walsh a,b
a
Department of Science, Institute of Technology Tallaght, Dublin 24, Ireland
b
Centre for Applied Science and Health, Institute of Technology Tallaght, Dublin 24, Ireland
c
The Inorganic Pharmaceutical and Biomimetic Research Group, Focas Research Institute, Dublin Institute of Technology, Camden Row, Dublin 8, Ireland
d
Institute of General and Inorganic Chemistry, Bulgarian Academy of Sciences, 11 Acad. G. Bonchev Str., Sofia, Bulgaria
e
Chemistry Department, National University of Ireland, Maynooth Co. Kildare, Ireland

a r t i c l e i n f o a b s t r a c t

Article history: Novel Ni(II), Co(II), Zn(II) and Mn(II) complexes of coumarin-3-carboxylic acid (HCCA) were studied at
Received 14 March 2011 experimental and theoretical levels. The complexes were characterised by elemental analyses, FT-IR,
Received in revised form 1 September 2011 1
H NMR, 13 C NMR and UV–Vis spectroscopy and by magnetic susceptibility measurements. The binding
Accepted 15 September 2011
modes of the ligand and the spin states of the metal complexes were established by means of molecu-
lar modelling of the complexes studied and calculation of their IR, NMR and absorption spectra at
Keywords:
DFT(TDDFT)/B3LYP level. The experimental and calculated data verified high spin Ni(II), Co(II) and Mn(II)
Coumarin
complexes and a bidentate binding through the carboxylic oxygen atoms (CCA2). The model calculations
Metal complexes
DFT modelling
predicted pseudo octahedral trans-[M(CCA2)2 (H2 O)2 ] structures for the Zn(II), Ni(II) and Co(II) complexes
TDDFT and a binuclear [Mn2 (CCA2)4 (H2 O)2 ] structure. Experimental and calculated 1 H, 13 C NMR, IR and UV–Vis
IR data were used to distinguish the two possible bidentate binding modes (CCA1 and CCA2) as well as
NMR mononuclear and binuclear structures of the metal complexes.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction The coordination chemistry and biological activity of coumarin


derivatives have also been of interest to our group [9,14–16]. In
Coumarin (2H-1-benzopyran-2-one) derivatives are oxygen- the past, we have reported a series of Cu(II) and Ag(I) coumarin
containing heterocyclic compounds that are present in a large carboxylates, which showed promising results in a number of
number of natural products [1,2]. The diversity of applications biological assays [17–19]. For a number of these complexes the
for coumarin-based compounds has resulted in well established isolation of crystals suitable for X-ray analysis often remained
general methods for their syntheses and still attracts organic and unsuccessful, and hence their coordination modes were pro-
medicinal chemists towards improvements and modifications of posed based on experimental data other than X-ray diffraction
already existing methods [3–7]. It is also well known, that coumarin [16,20–23]. As a continuation of this work, a number of coumarin-
derivatives can yield a wide variety of metal complexes with dif- 3-carboxylic acid-derived complexes containing ions other than
ferent coordination modes, spectroscopic properties, and potential Cu(II) and Ag(I) were synthesised. Bidentate coordination through
applications. In particular, the complexation of coumarin-derived the lactone carbonylic oxygen and the deprotonated carboxylic
ligands to various d-block metal ions has been recognised as a oxygen atoms (CCA1) has been previously proposed for a num-
promising route towards development of new therapeutic agents ber of lanthanide cations (La(III), Ce(III), Nd(III)) with CCA [21,22].
[8–13]. Model calculations of this coordination type, including geome-
try optimization, IR spectra simulations and detailed vibrational
analysis proved the CCA1 binding for the lanthanide ions. An alter-
native coordination of CCA, through the carboxylic oxygen atoms,
∗ Corresponding author at: Department of Science, Institute of Technology Tal- CCA2, was predicted for Cu(II) complexes of HCCA [23]. Therefore,
laght, Dublin 24, Ireland. Tel.: +353 01 404 2889. in addition to the available experimental data, DFT modelling of the
∗∗ Corresponding author. Tel.: +359 2 9792592.
E-mail addresses: Bernie.Creaven@ittdublin.ie (B.S. Creaven),
complexes was performed. These efforts allowed a confirmation of
ivelina@svr.igic.bas.bg (I. Georgieva). the binding mode in the HCCA-derived complexes.

1386-1425/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.saa.2011.09.041
276 B.S. Creaven et al. / Spectrochimica Acta Part A 84 (2011) 275–285

2. Materials and methods shifts (ıi =  TMS −  i ) for C and H atoms by referring to the stan-
dard compounds, tetramethylsilane (TMS). The referred compound
2.1. Physical measurements TMS was also calculated in DMSO at the same level of theory.
To simulate UV–Vis absorption spectra of Zn(II), Co(II), Ni(II) and
All chemicals were purchased from Sigma–Aldrich and used Mn(II) complexes of CCA, 50 lowest singlet excited states were cal-
without purification. Infrared spectra were recorded in the region culated in DMSO by means of TDDFT method and B3LYP/Basis1
of 4000 cm−1 to 400 cm−1 on Nicolet Impact 410 Fourier-Transform level [31]. In spite of difficulties of TDDFT approach in treating
Infrared (FTIR) spectrophotometer using Omnic software. All charge-transfer transitions properly [32] it remains a computation-
infrared spectra were run as KBr discs. Melting point values were ally simple and efficient method able to treat practical problems in
recorded on a Stuart scientific SMP1 melting point apparatus. a reasonable time scale at low cost as compared to highly corre-
Values were taken up to 320 ◦ C. UV–Vis spectra were recorded lated ab initio methods [33]. TDDFT calculations have proven to be
on a Hitachi U-2001 spectrophotometer in DMSO (spectopho- quite reliable in describing also the electronic spectra of complexes
tometric grade) in the region of 260–900 nm. All samples were containing transition metals in solution [34]. For comparison,
dissolved in d6 -DMSO. Signal assignments were made using stan- the vertical excitation energies of HCCA and its Zn(II) complex
dard techniques including DEPT/COSY and CHSHiFt experiments. were calculated at PBE1PBE level [35], recently established as a
Microanalytical data were provided by the Microanalytical Labora- more reliable hybrid functional for description of high-energetic
tory, National University of Ireland, Belfield, Dublin 4, Ireland. Room excitations of coumarins [36,37]. NMR and UV–Vis spectra calcu-
temperature magnetic susceptibility measurements were carried lations in DMSO were performed for the optimized geometries
out using a Johnson Matthey magnetic balance in NUI, Maynooth, in the same solvent. A non-equilibrium implementation of the
Co. Kildare, Ireland. Hg[Co(SCN)4 ] was used as the calibration stan- Polarisable Continuum Model (PCM) using the integral equation
dard. formalism variant (IEFPCM) was applied to simulate the solvent
effects [38–40].
2.2. Computational details

3. Results and discussion


Molecular structure, vibrational (IR), NMR and absorption spec-
tra calculations of M(CCA)2 (H2 O)2 (M = Zn(II), Co(II), Ni(II), Mn(II))
3.1. Preparation of complexes
and Mn2 (CCA)4 (H2 O)2 were carried out with DFT/TDDFT method
using Gaussian09 program [24]. The data were compared with
Commercially available coumarin-3-carboxylic acid (HCCA)
the corresponding results for the isolated neutral ligand (HCCA)
(Fig. 1a) was complexed to Zn(II), Co(II), Ni(II) and Mn(II) salts
at the same level of theory. Basis set 6-31G(d) (split valence
resulting in the formation of the corresponding complexes,
plus polarization basis set) was applied for the main group ele-
namely Bis[coumarin-3-carboxylato]zinc(II) dihydrate (1), CCA-
ments: for hydrogen atoms (4s)/[2s], for carbon, and oxygen atoms
Zn; Bis[coumarin-3-carboxylato]cobalt(II) dihydrate (2), CCA-Co;
(10s4p1d)/[3s2p1d]. To give a better description of the wave func-
Bis[coumarin-3-carboxylato]nickel(II) dihydrate (3), CCA-Ni; and
tions in the intermolecular region and hence, to satisfy estimation
Tetrakis[coumarin-3-carboxylato]dimanganese(II) dihydrate (4),
of the M-CCA and M-OH2 interactions, diffuse functions were
CCA-Mn using a modified procedure reported previously by Kar-
added to the standard basis set (one s and one p set) for the
aliota et al. [23]. Details are given in the supplementary section.
O atoms as well as for the C1, C2 and C3 atoms (Fig. 1). For
The addition of 0.02 M NaOH (5 ml) prior to the addition of the
Zn, Co and Ni, the Watchers’ primitive set (14s9p5d) [25] was
relevant salt, facilitated deprotonation of the carboxylic acid. Com-
employed supplemented with one s, two p and one d diffuse func-
plexes were then prepared by addition of aqueous solutions of
tions [26], and one f polarization function [27]. The final basis set
the relevant metal salt to DMF solutions of coumarin-3-carboxylic
6-311+G(d) is of the form (15s11p6d1f)/[10s7p4d1f]. Hereafter,
acid under continuous stirring and refluxing. All the complexes
the combined basis set for the metal complex molecule will be
were easily isolated from the hot reaction mixture by filtration
referred as Basis1. Geometry optimization and absorption spectra
as the solubility of all the products was significantly lower than
calculations of two model complexes of Mn(II), Mn(CCA2)2 (H2 O)2
that of free HCCA. Although the formation of binuclear structures
and Mn2 (CCA2)4 (H2 O)2 , were performed with contracted 6-31G(d)
was considered, the elemental analysis data for the zinc, cobalt
basis set to prevent time consuming calculations. The computations
and nickel complexes correlated better with a mononuclear struc-
using Basis1 and 6-31G(d) basis sets were carried out with the Lee,
ture, (Table 1). Only the results for the Mn(II) complex 4 were
Yang and Parr correlation functional (LYP) [28] combined with the
consistent with the presence of two metal ions. Microanalytical
Becke’s non-local three-parameter exchange functional, (B3) [29].
data supported the presence of water molecules as additional lig-
The adequacy of the B3LYP method for prediction of conforma-
ands.
tional behaviour, geometrical parameters and vibrational spectrum
of HCCA was proved in our recent investigations [21,22]. Full geom-
etry optimizations were carried out without symmetry constraints. 3.2. Magnetic susceptibility and spin states of the metal
The minima on the potential energy surfaces were qualified by complexes
the absence of negative eigenvalues in the diagonalized Hessian
matrix, giving imaginary normal vibrational mode. The approxi- As expected, the zinc complex, 1, revealed diamagnetic prop-
mate assignment of the calculated frequencies to the molecular erties and did not exhibit a spin moment. The effective magnetic
vibrational modes was performed using the atom movements in moment obtained for CCA-Co (2) was 5.47 BM consistent with a
Cartesian coordinate calculated with the Gaussian program. The high spin octahedral complex. This value is slightly larger than the
calculated atom movements were further used as an input of the characteristic experimental eff values for high spin Co(II) com-
ChemCraft program which animates them and gives the possi- plexes (4.1–5.2 BM) and the calculated spin-only magnetic moment
bility for a visual inspection of each molecular mode [30]. The (so 3.87 BM) [41]. The higher than expected experimental eff
1 H and 13 C NMR chemical shift calculations of HCCA and model value might be as a result of a considerable orbital moment s
M(CCA)2 (H2 O)2 and Mn2 (CCA)4 (H2 O)2 complexes were performed for 2 which contributes to the overall magnetic moment [42]. eff
in DMSO at B3LYP/6-31+G(d,p) level. The absolute isotropic mag- values of approximately 5.0 BM are known to be characteristic of
netic shielding constants ( i ) were used to obtain the chemical high spin octahedral Co(II) complexes [43]. For CCA-Ni (3) the eff
B.S. Creaven et al. / Spectrochimica Acta Part A 84 (2011) 275–285 277

Fig. 1. DFT/B3LYP/Basis1 optimized structures of: (a) neutral ligand HCCA, (b) trans-M(CCA1)2 (H2 O)2 and (c) trans-M(CCA2)2 (H2 O)2 , [M = Zn(II), Co(II), Ni(II) and Mn(II)].

Table 1
Physical data for metal(II) complexes of HCCA ligand 1–4.

Complex Calc (found) [%] Mw [g/mol] Yield [%] Empirical formula m.p. [◦ C] eff [BM]

C H M*

CCA-Zn (1) 50.07 (50.11) 2.94 (2.92) 13.63 (13.96) 479.71 73 C20 H10 O8 Zn·2H2 O 206–208 –
CCA-Co (2) 50.76 (50.66) 2.98 (3.02) 12.45 (–) 473.25 87 C20 H10 O8 Co·2H2 O 254–256 5.47
CCA-Ni (3) 50.78 (50.64) 2.98 (3.01) 12.41 (12.78) 473.01 85 C20 H10 O8 Ni·2H2 O >300 3.42
CCA-Mn (4) 54.07 (54.16) 2.95 (2.89) 12.37 (–) 888.50 81 C40 H20 O16 Mn2 ·2H2 O >300 7.38

* = Metal, – = not determined.


278 B.S. Creaven et al. / Spectrochimica Acta Part A 84 (2011) 275–285

value was 3.42 BM which correlates well with the typical eff val- to those obtained experimentally. Comparisons of structures with
ues for high spin Ni(II) octahedral complexes (2.9–3.4 BM) [41] CCA1 binding (trans-[M(CCA1)2 (H2 O)2 ]) and their spectroscopic
(calc. so 2.83 BM) and it is similar to values reported for other properties were also carried out. These efforts allowed determi-
mononuclear, deformed octahedral Ni(II) complexes containing nation of the spectroscopic properties specific to CCA1 and CCA2
carboxylate-substituted ligands (3.04–3.28 BM) [44,45]. CCA-Mn binding modes in the metal complexes 1–3.
(4) exhibited an effective magnetic moment value of 7.38 BM and The elemental analysis and effective magnetic moment value
was consistent with high-spin octahedral Mn(II) complexes. The of 7.38 BM suggested a binuclear structure for the Mn(II) com-
value is much higher than the values characteristic for high-spin plex of CCA. At the same time, the IR spectrum of the Mn(II)
mononuclear structures (∼5.9 BM, calc. so 5.92 BM) [41] and is complex is very similar to those of the mononuclear Zn(II), Co(II)
indicative of a high-spin binuclear complex. A similar value was and Ni(II) complexes of CCA. Therefore, both mononuclear trans-
reported for a binuclear Mn(II) structure with bridging carboxylate [Mn(CCA2)2 (H2 O)2 ] and binuclear [Mn2 (CCA2)4 (H2 O)2 ] complexes
ligands reported by Kiskin et al., eff = 7.0 BM at a temperature of were considered and their IR, NMR and absorption spectra were cal-
300 K [46]. culated and compared with the available spectroscopic data. Both
model structures possess minima in their potential energy surface
3.3. Molecular modelling of Zn(II), Co(II), Ni(II) and Mn(II) and both are stabilized via H-bonds formed between the coor-
complexes of CCA in the gas phase dinated water hydrogens and coumarin lactone-carbonyl atoms
(Figs. 1c and 2).
Molecular electrostatic potential calculations of the active
deprotonated form of the ligand (CCA) suggested two possible 3.4. Experimental and theoretical NMR spectra
sites of bidentate binding to the metal ion: first through the car-
bonyl and deprotonated carboxylic oxygens (CCA1), and second The experimental and calculated 1 H and 13 C NMR chemical
through both the carboxylic oxygens (CCA2) [21]. All possible spin shifts of HCCA and its [Zn(CCA)2 (H2 O)2 ] complex in DMSO are
states of Co(II), Ni(II) and Mn(II) complexes were calculated for the presented in Table 2. The trend in the experimental chemical
model [MCCA1]+ and [MCCA2]+ species and their relative electronic shift in HCCA and its Zn(II) complex was followed and compared
energies are given in Table S1 (Supplementary material). The cal- with the calculated ones for model HCCA, trans-[Zn(CCA1)2 (H2 O)2 ]
culated energies of the models studied predicted a stabilization of and trans-[Zn(CCA2)2 (H2 O)2 ]. A comparative NMR analysis gives
the [MCCA2]+ species (bidentate binding through the carboxylic additional evidence for the complex formation and the ligand coor-
oxygens) in the high spin states, namely a triplet state for Ni(II), a dination mode.
quartet state for Co(II), and a sextet state for the Mn(II) complex. The measured 1 H NMR spectrum of HCCA and its correspon-
The results obtained are in very good agreement with the metal ding Zn(II) complex revealed five signals in the aromatic region.
spin state predictions based on the effective magnetic moments Four of these signals were assigned to the benzene ring of the
found (eff ). Furthermore, all calculations were performed for the coumarin moiety. The fifth peak in this region was attributed to the
high spin state complexes. vinyl H4 proton. Upon complexation the positioning of this peak
The CCA1 and CCA2 coordination modes were also exami- shifted upfield, from 8.73 to 8.47 ppm. The 1 H NMR calculations
ned for the [M(CCA)2 (H2 O)2 ] complexes as suggested by the revealed that the positioning of the H4 signal varied for trans-
elemental analysis, M = Zn(II), Co(II) and Ni(II) (Fig. 1b and c). [Zn(CCA1)2 (H2 O)2 ] and trans-[Zn(CCA2)2 (H2 O)2 ] complexes. The
Both cis- and trans-CCA isomers were considered. Geometry value of 9.04 ppm calculated for the free HCCA ligand, changed to
optimizations of the model complexes showed several energy 7.97 ppm on simulation of the trans-[Zn(CCA2)2 (H2 O)2 ] formation,
minima structures with difference of 2–3 kcal/mol. The relative while for the trans-[Zn(CCA1)2 (H2 O)2 ] the positioning of H4 was
energies are given in Table S2 (Supplementary material). The opti- calculated as 9.47 ppm. The calculated upfield shift of H4 for trans-
mized [M(CCA1)2 (H2 O)2 ] complexes (bidentate CCA coordination [Zn(CCA2)2 (H2 O)2 ] complex was in agreement with the acquired
through carboxylic and lactone-carbonyl oxygens) show pseudo data. This might be evidence for the actual coordination mode of
octahedral structures and the trans-CCA1 position is the preferred the complex.
one (Fig. 1b). The bidentate CCA2 binding via both carboxylic oxy- The calculated chemical shift value for the carboxylic proton
gens leads to more stable structures compared to that of CCA1. was 12.96 ppm. This signal was not observed in the acquired HCCA
The trans-[M(CCA2)2 (H2 O)2 ] complexes reveal deformed octahe- spectrum, probably as a result of rapid exchange of the carboxylic
dral structures stabilized by two Hw · · ·O hydrogen bonds formed proton with protons of residual water in the DMSO solvent. All
between the coordinated water hydrogens and coumarin carbonyl the remaining signals either did not change their positioning upon
atoms (Fig. 1c). In contrast to the trans-[M(CCA1)2 (H2 O)2 ] struc- complexation or the changes in their shifts were minimal.
ture the aromatic ring of the trans-[M(CCA2)2 (H2 O)2 ] structure The 13 C NMR data for HCCA and its Zn(II) complex were
is perpendicular to the C3O3O4M plane. The model calculations also compared to their corresponding calculated values. In the
predicted another energy minimum structure stabilized by four HCCA spectrum, the signal at 148.3 ppm was assigned to the
intramolecular O· · ·Hw hydrogen bonds and tending to tetrahe- vinyl C4 carbon. The positioning of this peak seemed very sen-
dral metal coordination with monodentate CCA binding via the sitive to environmental changes and might be particularly useful
carboxylic oxygen; however, such a tetrahedral coordination is in in the determination of the coordination mode of the complex.
conflict with the effective magnetic moment values for Ni(II) and The acquired data showed a migration of the C4 signal upon
Co(II) and their predicted octahedral coordination modes. More- complexation to 144.7 ppm. According to the calculations this
over, the calculated IR spectrum of the tetrahedral monodentate result indicated the CCA2 coordination of the complex (ligand
[M(CCA)2 (H2 O)2 ] complex shows an intense band at 1313 cm−1 due calc. 151.5 ppm → complex calc. 137.5 ppm). Calculated values for
to stretching ␯(C3O3) vibration, which is not observed in the exper- CCA1 binding mode resulted in a downfield shift of the C4 sig-
imental IR spectrum of the CCA-M(II) complexes. Therefore, for nal (ligand calc. 151.5 ppm → complex calc. 152.9 ppm). Another
Zn(II), Ni(II) and Co(II) complexes we accepted bidentate binding of significant downfield shift upon complexation was observed for
CCA via the carboxylic oxygens in the trans position and octahedral the C2 carbon (116.2 ppm → 124.4 ppm). The corresponding calcu-
complex structures, [M(CCA2)2 (H2 O)2 ] (Fig. 1c). In order to con- lations supported the CCA2 coordination mode. In the 13 C NMR
firm this hypothesis, the IR, NMR and UV–Vis absorption spectra at spectrum of HCCA, the C1 signal was visible at 164.1 ppm and the
the optimized octahedral structures were calculated and compared C3 signal – at 156 ppm. Upon complexation, the C1 signal was
B.S. Creaven et al. / Spectrochimica Acta Part A 84 (2011) 275–285 279

Fig. 2. B3LYP/6-31G(d) optimized structure of Mn2 (CCA2)4 (H2 O)2 .

shifted upfield (157.8 ppm), whereas the C3 signal was shifted the Zn(II) complex and they are given in Table S3 (Supplemen-
downfield (167.4 ppm). Changes in the positioning of these sig- tary material). Comparison of the calculated 13 C NMR data for
nals correlated with their corresponding calculated results for the [Mn(CCA2)2 (H2 O)2 ] and [Mn2 (CCA2)4 (H2 O)2 ] complexes indicated
trans-[Zn(CCA2)2 (H2 O)2 ]. Overall, the combined acquisition and a larger downfield positioning of C2 and C3 atoms and a larger
calculated NMR results for HCCA and its Zn(II) complex gave a upfield shift of C4 atom for the first complex, Table S4 (Supple-
strong indication of the CCA2 coordination mode of the complex. mentary material). The H4 signal did not change in the binuclear
The calculated NMR results for trans-[Ni(CCA1/CCA2)2 (H2 O)2 ] Mn(II) complex, whereas it moved to higher field in mononuclear
and trans-[Co(CCA1/CCA2)2 (H2 O)2 ] complexes showed similar [Mn(CCA2)2 (H2 O)2 ] complex. Based on these results it was possible
trends in the chemical shifts upon complexation as those of to distinguish between mono- and binuclear Mn(II) complex.

Table 2
Calculated 1 H and 13 C NMR chemical shifts (ppm) in DMSO at B3LYP/6-31+G(d,p) level for HCCA ligand and its Zn(II) complex compared to the experimental data in DMSO.

Atom HCCA Zn(CCA)2 (H2 O)2

Exp Calc Exp CCA2 calc CCA1 calc


a
H (C4) 8.73 9.04 8.47 7.97 9.47
H (C6) 7.88d 8.01 7.84d 7.81 8.04
H (C9) 7.88d 7.79 7.84d 7.88 7.82
H (C7) 7.71t 8.22 7.64t 7.67 7.75
H (C8) 7.71t 7.73 7.64t 7.92 8.06
H(O3) – 12.96 – – –
H(O)w – 2.80 2.57, 2.46
H(O)w – 6.54 (HB)
C1 164.1 162.6 157.8 160.4 163.2
C3 156.7 161.9 167.4 175.8 163.4
C10 154.6 153.5 154.0 153.7 154.3
C4 148.3 151.5 144.7 137.5 152.9
C8 134.4 134.8 132.7 130.8 133.7
C6 130.3 130.1 129.3 129.7 129.3
C7 124.9 123.7 123.4 123.4 123.3
C5 118.4 119.5 118.5 117.6 119.1
C9 117.4 114.6 115.8 114.9 113.6
C2 116.2 112.4 124.4 124.7 116.4
a
Atom numbering is given in Fig. 1.
280 B.S. Creaven et al. / Spectrochimica Acta Part A 84 (2011) 275–285

Fig. 3. TDDFT/B3LYP/Basis1 absorption spectra of free HCCA and trans-M(CCA2)2 (H2 O)2 complexes in DMSO, M = Ni(II), Co(II) and Zn(II).
B.S. Creaven et al. / Spectrochimica Acta Part A 84 (2011) 275–285 281

Fig. 4. Calculated absorption spectra of Mn2 (CCA2)4 (H2 O)2 and Mn(CCA2)2 (H2 O)2 complexes at B3LYP/6-31G(d) level.

The 1 H and 13 C NMR spectra for Ni(II), Co(II) and Mn(II) com- in DMSO solution. Such underestimation of the transition ener-
plexes were not acquired due to the strong paramagnetic character gies at B3LYP level is known from previous studies of coumarins
of these compounds. [36,37,47]. TDDFT/PBE1PBE/Basis1 calculations of HCCA and trans-
[Zn(CCA2)2 (H2 O)2 ] in DMSO showed higher excitation energies by
3.5. Experimental and theoretical UV–Vis spectra 0.1–0.15 eV as compared to the B3LYP data, in line with recent
theoretical studies [36,37]. In case of HCCA, PBE1PBE vertical excita-
The measured UV–Vis spectra of HCCA and its Ni(II), Co(II) and tion energies better approach the experiment (than those at B3LYP
Zn(II) complexes were analyzed on the basis of TDDFT calcula- level), however the second 1 A (␲␲*) excited state remains in being
tions of the vertical excitation energies in DMSO (Supplementary lower by ∼0.2 eV. In the case of the Zn(II) complex however, the
material, Table S5). The calculated absorption spectra and the PBE1PBE results more deviate from the experimental excitation
character of the transitions for HCCA and trans-[M(CCA2)2 (H2 O)2 ] energies (as compared to the B3LYP data). Hence, the B3LYP/Basis1
complexes (M = Ni(II), Co(II) and Zn(II)) are given in Fig. 3. The excitation energies of the metal complexes of HCCA are discussed
experimental absorption spectrum of HCCA revealed two bands below.
with maxima at 329 nm (low intensity) and 285 nm respectively. For the trans-[Ni(CCA2)2 (H2 O)2 ] complex, the calculated verti-
The calculated absorption maxima at 329 nm and 308 nm corre- cal excitation energies at 319 nm and 292 nm with large oscillator
late with the experimental data and correspond to the vertical strengths are consistent with the observed absorption maxima at
excitation energies of the first 1 A (␲␲*) and second 1 A (␲␲*) 311 nm and 297 nm. This result confirmed the structure suspected.
excited states, respectively (Fig. 3). According to the calculations According to Mulliken population analysis, the excitation related
␲ → ␲* transitions with large oscillator strengths are expected to the first absorption band at 311 (exp.)/319 (calc.) nm involves
also at 221 nm and 209 nm in the UV–Vis spectrum of HCCA transition from CCA(␲) orbital to a mixed orbital with Ni(dxz ) and
(Supplementary material, Table S5). CCA(␲*) contributions. Therefore, this band has ligand-to-mixed
The calculated B3LYP energy of the second 1 A (␲␲*) excited metal/ligand charge transfer character and it appears characteris-
state of HCCA is lower by 0.32 eV than the experimental one tic for the [Ni(CCA2)2 (H2 O)2 ] complex (Fig. 3). The second band at
282 B.S. Creaven et al. / Spectrochimica Acta Part A 84 (2011) 275–285

Fig. 5. Calculated (scaled by factor 0.951) IR spectra of HCCA, Zn(CCA1)2 (H2 O)2 and Zn(CCA2)2 (H2 O)2 complexes at B3LYP/Basis1 level (a) and of Mn2 (CCA2)4 (H2 O)2 and
Mn(CCA2)2 (H2 O)2 at B3LYP/6-31G(d) level.

297 (exp.)/292 (calc.) nm reveals CCA(␲) → CCA(␲*) character. The complex of CCA revealed metal-to-ligand charge transfer character,
d-d transitions are calculated in the range 890–500 nm with very it is due to a transition from a mixed Mn(dxz )/CCA(␲) orbital to a
low intensity. CCA(␲*) orbital. Unfortunately, bands at these wavelengths were
For the complex trans-[Co(CCA2)2 (H2 O)2 ] the calculated not observed in the experimental UV–Vis spectrum.
absorption bands at 317, 305, 295, 292 nm respectively, correlate
well with the observed UV–Vis bands at 320, 308, 297, 294 nm
(Fig. 3 and Table S5). According to the population analysis of 3.6. Experimental and theoretical IR spectra
the model Co(II) complex, bands at 317, 305 and 295 nm are
attributed to metal–ligand charge transfer transitions. The first Geometry optimization and IR spectrum calculations were per-
band represents a mixed Co(dxz )/CCA(␲) orbital to a CCA(␲*) orbital formed for the trans-[Zn(CCA2)2 (H2 O)2 ] complex. The calculated
transition, while the second and the third ones represent a CCA(␲) and scaled by factor 0.951 IR spectra in the range 200–1800 cm−1 of
to a Co(dyz )/CCA(␲*) orbital transitions. Bands at 305 nm and HCCA, trans-[Zn(CCA1)2 (H2 O)2 ] and trans-[Zn(CCA2)2 (H2 O)2 ] are
295 nm reveal significant intraligand charge transfer contribution, shown in Fig. 5a. Selected experimental and calculated IR data and
CCA(␲) → CCA(␲*). assignments of HCCA and both trans-[Zn(CCA)2 (H2 O)2 ] model com-
TDDFT calculations of the vertical excitation energies for plexes are given in Table 3.
trans-[Zn(CCA2)2 (H2 O)2 ] complex predict bands at 309, 298, and The strong IR band observed at 3200 cm−1 in the IR spectrum
292 nm with CCA(␲) → CCA(␲*) transition character (Fig. 3). These of CCA-Zn was assigned to a ␯(OH) vibration. Different ␯(OH) band
bands are related to the observed absorption bands at 322, positioning was calculated for CCA1 and CCA2 coordination modes
297, and 291 nm respectively. d-Contribution for these transi- and this finding can be used to distinguish the binding mode in
tions was not calculated. The intense band at 202 nm reveals the metal complexes of CCA. The calculated geometry of trans-
CCA(␲) → Zn(s)/CCA(␲*) character. [Zn(CCA2)2 (H2 O)2 ] revealed H-bond formation Hw · · ·O2 (Fig. 1c)
The absorption spectra of the two model Mn(II) complexes resulting in elongation of O–H bond lengths. Therefore, ␯(OH) fre-
were simulated by TDDFT calculations (Fig. 4 and Table S6). For quency was expected to shift to lower frequencies compared to
the mononuclear complex of Mn(II), [Mn(CCA2)2 (H2 O)2 ], a band its positioning in [Zn(CCA1)2 (H2 O)2 ] complex where intramolec-
at 302 nm with CCA(␲) → CCA(␲*) character was calculated. For ular H-bonding was not present (Fig. 1b). Thus, the calculated
[Mn2 (CCA2)4 (H2 O)2 ] a complex high intense absorption band at ␯as (OH) for trans-[Zn(CCA2)2 (H2 O)2 ] was obtained at lower fre-
321 nm was calculated and assigned to a ligand-to-ligand charge quencies (3316 cm−1 ) in better agreement with the experimental
transfer transition. The 321 nm value better corresponds to the data than in trans-[Zn(CCA1)2 (H2 O)2 ] (3452 cm−1 ). The experimen-
band observed at 327 nm and supports the formation of a binuclear tal IR spectra of the Ln(III)-CCA complexes [22], where the CCA1
CCA complex of Mn(II). The second high intensity band calculated at binding is operative, showed a ␯(OH) band at ∼ 3430 cm−1 , whereas
295 nm for the binuclear and at 285 nm for the mononuclear Mn(II) the experimental IR spectra of the M(II)-CCA complexes (M = Zn, Co,
B.S. Creaven et al. / Spectrochimica Acta Part A 84 (2011) 275–285 283

Table 3
Selected experimental and calculateda IR frequencies (in cm−1 ) and IR intensities (I, in km mol−1 ) of HCCA, Zn(CCA2)2 (H2 O)2 and Zn(CCA1)2 (H2 O)2 complexes.

HCCA Zn compl Zn(CCA2)2 (H2 O)2 Assignment Zn(CCA1)2 (H2 O)2

Exp. Calc. Exp. Calc. Calc.

Freq. I Freq. I Freq. I

Overlap 3607 275 ␯as (OH)w 3614 166


3607 0 ␯s (OH)w 3614 7
– – – 3200s 3323 1931 ␯as (OH)w 3469 176
3323 25 ␯s (OH)w 3468 5
3056s 3243 407 – – – ␯(O–H)coum – –
1745vs 1729 447 – – – ␯(C3 O4)fr carbox 1649 779
1682vs 1657 552 1667s 1663 1760 ␯(C1 O2)carbonyl 1576 972
1616s 1633 193 ␦(OHO) 1620 421
1609s 1583 264 1600m,d 1603 81 ␯(CC) + ␦(CCH) 1607 253
1568s 1533 180 1585s 1583 236 ␯(CC) + ␦(CCH) 1569 52
– – – 1561s 1549 596 ␯as (CO2 ) + ␦(CCO) – –
1489 1459 10 1457s 1434 111 ␯(CC) + ␦(CCH) 1529 388
1458 14
1422 43
1452s 1374 254 – – ␦(COH) + ␯(O· · ·H) – –
1421sh
– – – 1396s 1389 482 ␯s (CO2 ) + ␯(CC) – –
1373s 1335 23 – – ␯(CC) + ␦(CCH) 1328 15
– – 1288m – – ␯(C3O3) 1282 720
1298w 1318 1 1288m 1331 32 ␦(CCH) + ␯(CC) – –
1244w 1260 21 1259m 1250 108 ␯(CC) + ␦(CCH), 1247 361
1226 0 1221m 1214 66 ␯(CC) + ␦(CCH) 1217 153
1227m 1194 56 – – – ␯(C10O1) 1188 8
1206m 1176 137 1184m 1146 97 ␦(CCH) + ␯(C–Olac ) 1138 130
1164m 1136 16 1153m 1121 31 ␦(CCH) + ␯(C–Olac ) 1115 130
1121m 1106 16 1132m 1095 21 ␦(CCH) 1095 18
1041s 986 54 1046m 1014 168 ␯(C1O1) 995 44
989s 959 13 1024w 1005 19 ␦(CCH)o.o.p. 1005 13
963w 950 3 974w 945 0 ␦(CCH)o.o.p. 991 55
938m 914 1 929w 921 32 ␦(CCH)o.o.p. 955 20
924m 887 5 879m 889 15 ␦(CCC)ip – –
882m 841 3 – – – ␦(CCH)o.o.p. – –
832m 791 74 – – – ␦(COH)o.o.p. – –
810m 795 312 ␦(OHO) ␦(ZnOHHB ) – –
801m 751 50 779s 767 129 t(OCCO2 )ip /Zn-O3 753 128
741m 749 166 t(CCCO2 )
770m 739 55 711w 736 121 ␦(CCH)o.o.p.
744w 723 5 646s 696 14 ␦(CCC)ip
623w 590 65 ␦(HOH) 593 424
571 214
572 79 Zn-Oc
593m 546 243 ␦ (ZnOH)
587w 619 12 ␯(O· · ·H)
574w 566 9 577m 538 28 Ringip
473m 521 30 ␦(CCO)ip /␦ (HOH) 472 519
470s 450 15 463m 504 8 ringip
381 243 ␯as (Zn-O), t(HOPtOH) 457 38
261 120 Zn-Ow 218 28
a
Calculations at DFT/B3LYP/6-31G(d) level, frequencies are scaled by factor 0.951.

Ni, Mn) with suggested CCA2 binding (this work) exhibited a ␯(OH) the Zn(II) complex spectrum is assigned to ␯s (CO2 ). The experi-
band at about 3200 cm−1 . mental peak at 1396 cm−1 shows splitting due to the presence
According to the calculations, the strong band observed at of two CCA ligands and this finding confirms the assignment.
1745 cm−1 in the IR spectrum of the free HCCA ligand is assigned It is worth mentioning that this band is missing in the calcu-
to the carboxylic ␯(C3 O4) mode (calc. 1729 cm−1 ). The calcu- lated IR spectrum of [Zn(CCA1)2 (H2 O)2 ] complex and therefore it
lations showed that the coordination of Zn(II) via the carboxylic could be considered as characteristic of CCA2 coordination. The
CO2 − group produces equalization of C3 O4 and C3–O3 bonds theoretically predicted ␯(OCO) = ␯as (CO2 ) − ␯s (CO2 ) difference is
(elongation of the double C3 O4 bond, ∼0.055 Å) and signifi- 160 cm−1 . Accepting the assignment of the experimental bands
cant lowering of the asymmetric ␯as (CO2 ) frequencies ( ∼180, mentioned above, the experimental ␯as (CO2 ) − ␯s (CO2 ) difference
209 cm−1 ). It should be noted that in HCCA, distinct ␯(C3 O4) is estimated of 165 cm−1 (1561 − 1396 cm−1 ) and it is in the range
and ␯(C3–O3) carboxylic vibrations are observed, whereas in trans- of the predicted values. Unfortunately, the magnitudes obtained
Zn(CCA2)2 (H2 O)2 they appear as ␯as (CO2 ) and ␯s (CO2 ) vibrations. are characteristic not only for bidentate binding of the CO2 − group,
The calculated ␯as (CO2 ) frequencies at 1549, 1520 cm−1 are related they are reported in the literature also for bridging coordination of
to the strong experimental band at 1560 cm−1 . Thus, the theoreti- the CO2 − group [48].
cally predicted shift of ca. 195 cm−1 approaches the experimental The vibrational behaviour of the lactone carbonylic moiety,
one, 184 cm−1 : ␯(C3 O4) 1745 cm−1 → ␯as (CO2 ) 1561 cm−1 . The ␯(C1 O2) was also considered. In the free HCCA ligand, the car-
␯s (CO2 ) frequency in the Zn(II) complex is calculated at 1390 bonylic group is included in intramolecular hydrogen bonding
and 1389 cm−1 and thereby the strong band at 1396 cm−1 in (Fig. 1a) and the ␯(C1 O2) frequency is slightly lowered: at
284 B.S. Creaven et al. / Spectrochimica Acta Part A 84 (2011) 275–285

1657 cm−1 (calc.)/1682 cm−1 (exp.). In the [Zn(CCA2)2 (H2 O)2 ] com- binuclear sextet spin state Mn(II) complex, [Mn2 (CCA)4 (H2 O)2 ],
plex, the lactone carbonylic group does not bind to the metal ion were suggested by elemental analysis and magnetic susceptibility
but it is included in hydrogen bonding, in this case with the coor- measurements. Energy calculations of model CCA-M(II) complexes
dinated H2 O molecules, Fig. 1c. Thus, the ␯(C1 O2) frequencies in with the possible spin states confirmed the stabilization of the
the complex should be close to the corresponding one of the ligand, high spin Ni(II), Co(II) and Mn(II) complexes. Two bidentate lig-
1657 cm−1 . This was the case of the Zn(II) complex, the calculated and binding modes, CCA1 (through carbonylic and deprotonated
␯(C1 O2) frequency is 1663 cm−1 (Table 3) and it indicates that the carboxylic oxygens) and CCA2 (through both carboxylic oxygens)
strong band observed at 1667 cm−1 in the complex spectrum is due were examined in the model complexes by means of a compari-
to the lactone carbonylic stretching mode. The ligand → complex son of their calculated and experimental IR, NMR and absorption
band shift is small, only 19 cm−1 and could not indicate coordina- spectra. The theoretical and experimental spectroscopic data are in
tion via the lactone carbonylic group (along with the deprotonated better agreement with the mononuclear trans-[M(CCA2)2 (H2 O)2 ]
carboxylic oxygen). models for Co(II), Zn(II), Ni(II) and binuclear [Mn2 (CCA2)4 (H2 O)2 ]
At the same time, for the alternative coordination mode, model. The bidentate CCA binding through the carboxylic oxygens
CCA1 via the carboxylic and carbonylic oxygens (Fig. 1b), the in trans-position and the suggested pseudo octahedral structures
␯(C3 O4) frequency of the uncoordinated group was predicted at were theoretically confirmed. The comparative 13 C and 1 H NMR
1649 cm−1 . As compared to the ligand, the calculated ␯(C3 O4) data analysis of HCCA and its Zn(II), Co(II), Ni(II) and Mn(II)
shift is 80 cm−1 (av.): 1729 cm−1 (ligand) → 1649 cm−1 (IR active, complexes showed that the 1 H and 13 C signal behaviour could
for [Zn(CCA1)2 (H2 O)2 ] complex). This shift value is much smaller distinguish the CCA1 and CCA2 bidentate binding manners as
in comparison with that calculated for the coordinated C3–O4 well as the mononuclear and binuclear Mn(II)complexes. ␯(OH)
group in CCA2 coordination (av. 195 cm−1 ). In agreement with frequencies of mononuclear trans-[M(CCA)2 (H2 O)2 ] and binuclear
the calculated IR spectra of the two Zn(II) complexes, the [Mn2 (CCA)4 (H2 O)2 ] complexes with CCA2-M binding are observed
band observed at 1667 cm−1 could be either due to the car- and calculated at lower frequencies ∼3200 cm−1 due to intra- or
bonylic stretching mode, ␯(C1 O2) of CCA2 coordination, or due intermolecular O· · ·Hw bonds. In CCA1-M(II) complexes ␯(OH)w
to the carboxylic stretching mode, ␯(C3 O4) of CCA1 coordi- appear at ∼3400 cm−1 . The strong bands observed at 1561 and
nation type. In the free CCA ligand, the double C1 O2 and 1389 cm−1 in the IR spectra of the CCA-M(II) complexes assigned
C3 O4 bonds differ and the ␯(C1 O2) vibration appears at to ␯as (CO2 ) and ␯s (CO2 ) modes explicitly reveal the CCA2 manner
higher frequencies with ∼90 cm−1 . In trans-[Zn(CCA1)2 (H2 O)2 ] and of binding in the M(II) complexes.
trans-[Zn(CCA2)2 (H2 O)2 ] (Fig. 1b,c), however the corresponding
uncoordinated C3 O4 and C1 O2 frequencies are very similar due Acknowledgements
to the H-bonding of C1 O2 group and its frequency lowering. Thus,
one may conclude that the IR spectral behaviour of the stretching This research was supported by the Technological Sector
carbonylic frequency of the ligand, ␯(C1 O2) and the metal com- Research Programme, Strand 1, under the European Social Fund.
plex ␯(C1 O2)/␯(C3 O4) is not characteristic and could not be used D.K. gratefully acknowledges the EMBARK fellowship for addi-
to elucidate the CCA coordination mode. tional funding. The 13 C NMR analysis for compound 1 was carried
According to the calculations and in agreement with other out with the kind assistance of Dr Brian Murray, using Avance
assignments in the literature, the band at 1616 cm−1 in the com- III 400 MHz, courtesy of Bruker Biospin, UK. I.G. and N.T. thank
plex spectrum is assigned to the characteristic bending modes of the National Science Fund of Bulgaria for the financial support
the coordinated water molecules [48]. (Grant Nos. DO-02-233/2008 and DCVP-02/2/2009). The calcula-
The experimental IR spectra of Ni(II), Co(II) and Zn(II) com- tions were performed on the computer cluster MADARA (Project
plexes are very similar and therefore we believe that the vibrational RNF01/0110/2008 NSF of Bulgaria).
assignments made for Zn(II) complex will be valid for Ni(II) and
Co(II) complexes of CCA as well. Appendix A. Supplementary data
The calculated IR spectra of [Mn(CCA2)2 (H2 O)2 ] and
[Mn2 (CCA2)4 (H2 O)2 ] are quite similar, Fig. 5b. Selective vibrational Supplementary data associated with this article can be found, in
frequencies are given in Table S7 (Supplementary material). At the online version, at doi:10.1016/j.saa.2011.09.041.
first sight, the carbonyl ␯(C1 O2) vibrations in the simulated IR
spectra above differ slightly (1693, 1732 cm−1 ). It should envisage
References
that in the complexes studied the carbonyl C1 O2 group is not
bound to Mn(II) but it is involved in hydrogen bonding. Hence, [1] H.E. Kleiner, X. Xia, J. Sonoda, J. Zhang, E. Pontius, J. Abey, R.M. Evans, D.D. Moore,
the carbonyl ␯(C1 O2) vibrations appear sensitive to the H-bond J. DiGiovanni, Toxicol. Appl. Pharm. 232 (2008) 337.
strengths, which are evaluated by DFT calculations. In conclusion, [2] T. Smyth, V.N. Ramachandran, W.F. Smyth, Int. J. Antimicrob. Agents 33 (2009)
421.
the calculations predict that the IR spectroscopic data could not [3] F. Borges, F. Roleira, M. Milhazes, L. Santana, E. Uriate, Curr. Med. Chem. 12
significantly help to distinguish mono- and binuclear Mn(II) com- (2005) 887.
plexes with the same ligand binding mode. The result obtained [4] D.S. Bose, A.P. Rudradas, M.H. Babu, Tetrahedron Lett. 43 (2002) 9195.
[5] B. Tyagi, M.K. Mishra, R.V. Jasra, J. Mol. Catal. 276 (2007) 47.
explains the similar IR spectra of the metal complexes studied [6] J. Jayashankaran, R.D.R.S. Manian, R. Raghunathan, Tetrahedron Lett. 47 (2006)
although the large effective magnetic moment for Mn(II) complex 2265.
of 7.38 BM suggests a binuclear complex. [7] D. Maes, S. Vervisch, S. Debenedetti, C. Davio, S. Mangelinck, N. Giubellina, N.
De Kimpe, Tetrahedron 61 (2005) 2505.
[8] M. Grazul, E. Budzisz, Coord. Chem. Rev. 253 (2009) 2588.
[9] B.S. Creaven, M. Devereux, D. Karcz, A. Kellett, M. McCann, A. Noble, M. Walsh,
4. Conclusions J. Inorg. Biochem. 103 (2009) 1196.
[10] A. Kulkarni, S.A. Patil, P.S. Badami, Eur. J. Med. Chem. 44 (2009) 2904.
[11] M.P. Sathisha, U.N. Shetti, V.K. Revankar, K.S.R. Pai, Eur. J. Med. Chem. 43 (2008)
Novel Zn(II), Co(II), Ni(II) and Mn(II) complexes of coumarin- 2338.
3-carboxylic acid were synthesized and characterised by FTIR, [12] K.B. Gudasi, M.S. Patil, R.S. Vadavi, Eur. J. Med. Chem. 43 (2008) 2436.
1 H NMR, 13 C NMR and UV–Vis spectroscopies and by DFT/B3LYP [13] G.B. Bagihalli, P.G. Avaji, S.A. Patil, P.S. Badami, Eur. J. Med. Chem. 43 (2008)
2639.
calculations. Mononuclear diamagnetic Zn(II), quartet spin state [14] B.S. Creaven, D.A. Egan, D. Karcz, K. Kavanagh, M. McCann, M. Mahon, A. Noble,
Co(II) and triplet spin state Ni(II) complexes, [M(CCA)2 (H2 O)2 ] and B. Thati, M. Walsh, J. Inorg. Biochem. 101 (2007) 1108.
B.S. Creaven et al. / Spectrochimica Acta Part A 84 (2011) 275–285 285

[15] B.S. Creaven, D.A. Egan, K. Kavanagh, M. McCann, M. Mahon, A. Noble, B. Thati, [26] P.J. Hay, J. Chem. Phys. 66 (1977) 4377.
M. Walsh, Polyhedron 24 (2005) 949. [27] K. Raghavachari, G.W. Trucks, J. Chem. Phys. 91 (1989) 1062.
[16] B.S. Creaven, D.A. Egan, K. Kavanagh, M. McCann, A. Noble, B. Thati, M. Walsh, [28] C. Lee, W. Yang, R.G. Parr, Phys. Rev. B37 (1988) 785.
Inorg. Chim. Acta 359 (2006) 3976. [29] A.D. Becke, J. Chem. Phys. 98 (1993) 5648.
[17] B. Thati, A. Noble, R. Rowan, B.S. Creaven, M. Walsh, M. McCann, D. Egan, K. [30] Available from http://www.chemcraftprog.com.
Kavanagh, Toxicol. in Vitro 21 (2007) 801. [31] R. Bauernschmitt, R. Ahlrichs, Chem. Phys. Lett. 256 (1996) 454.
[18] B. Thati, A. Noble, B.S. Creaven, M. Walsh, M. McCann, K. Kavanagh, M. Devereux, [32] A. Dreuw, J.L. Weisman, M. Head-Gordon, J. Chem. Phys. 119 (2003) 2943.
D.A. Egan, Cancer Lett. 250 (2007) 128. [33] C. Daniel, Coord. Chem. Rev. 238 (2003) 143.
[19] B. Thati, A. Noble, B.S. Creaven, M. Walsh, M. McCann, K. Kavanagh, M. Devereux, [34] V. Barone, F. Fabrizi de Biani, E. Ruiz, B. Sieklucka, J. Am. Chem. Soc. 123 (2001)
D.A. Egan, Cancer Lett. 248 (2007) 321. 10742.
[20] I. Georgieva, N. Trendafilova, B.S. Creaven, M. Walsh, A. Noble, M. McCann, [35] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 78 (1997) 1396.
Chem. Phys. 365 (2009) 69. [36] J. Preat, D. Jacquemin, V. Wathelet, J.-M. André, E.A. Perète, J. Phys. Chem. A 110
[21] Tz. Mihaylov, N. Trendafilova, I. Kostova, I. Georgieva, G. Bauer, Chem. Phys. (2006) 8144.
327 (2006) 209. [37] W. Zhao, Y. Ding, Q. Xia, J. Comp. Chem. 32 (3) (2011) 545.
[22] I. Georgieva, N. Trendafilova, W. Kiefer, V.K. Rastogi, I. Kostova, Vib. Spectrosc. [38] M.T. Cancès, B. Mennucci, J. Tomasi, J. Chem. Phys. 107 (1997) 3032.
44 (2007) 78. [39] M. Cossi, V. Barone, B. Mennucci, J. Tomasi, Chem. Phys. Lett. 286 (1998) 253.
[23] A. Karaliota, O. Kretsi, C. Tzougraki, J. Inorg. Biochem. 84 (2001) 33. [40] B. Mennucci, J. Tomasi, J. Chem. Phys. 106 (1997) 5151.
[24] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, [41] F.A. Cotton, G. Wilkinson, C.A. Murillo, M. Bochmann, Advanced Inorganic
G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato, Chemistry, 6th ed., John Wiley & Sons, New York, 1999.
X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M. Hada, [42] U. Muller, Inorganic Structural Chemistry, Wiley, Chichester, 1995.
M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, [43] B. Żurowska, A. Brzuszkiewicz, Polyhedron 27 (2008) 1623.
O. Kitao, H. Nakai, T. Vreven Jr., J.A. Montgomery, J.E. Peralta, F. Ogliaro, M. [44] K. Deka, N. Barooah, R.J. Sarma, J.B. Baruah, J. Mol. Struct. (Theochem) 827 (2007)
Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, R. Kobayashi, J. Nor- 44.
mand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar, J. Tomasi, M. Cossi, N. [45] P. Kopel, Z. Trávnícek, J. Marek, J. Mrozinski, Polyhedron 23 (2004) 1573.
Rega, N.J. Millam, M. Klene, J.E. Knox, J.B. Cross, V. Bakken, C. Adamo, J. Jaramillo, [46] M.A. Kiskin, I.G. Fomina, G.G. Aleksandrov, A.A. Sidorov, V.M. Novotortsev, Y.V.
R. Gomperts, E.E. Stratmann, O. Yazyev, A.J. Austin, R. Cammi, C. Pomelli, J.W. Rakitin, Z.V. Dobrokhotova, V.N. Ikorskii, Y.G. Shvedenkov, I.L. Eremenko, I.I.
Ochterski, R.L. Martin, K. Morokuma, V.K. Zakrzewski, G.A. Voth, P. Salvador, Moiseev, Inorg. Chem. Commun. 8 (2005) 89.
J.J. Dannenberg, S. Dapprich, A.D. Daniels, Ö.J. Farkas, B. Foresman, J.V. Ortiz, J. [47] I. Georgieva, N. Trendafilova, A. Aquino, H. Lischka, J. Phys. Chem. A 109 (2005)
Cioslowski, D.J. Fox, Gaussian 09, Revision A.1, Gaussian, Inc., Wallingford, CT, 11860.
2009. [48] K. Nakamoto, Infrared and Raman Spectra of Inorganic and Coordination Com-
[25] A.J.H. Watchers, J. Chem. Phys. 52 (1970) 1033. pounds, Part III, Wiley, New York, 1978.

You might also like